Open Access
3 November 2014 Super-resolution photoacoustic microscopy using photonic nanojets: a simulation study
Author Affiliations +
Abstract
Optical resolution photoacoustic microscopy (ORPAM) is important for various biomedical applications, such as the study of cellular structures, microcirculation systems, and tumor angiogenesis. However, the lateral resolution of a conventional ORPAM is limited by optical diffraction. In this work, we report a simulation study to achieve subdiffraction-limited super-resolution in ORPAM using microspheres. Laser radiation is focused through a microsphere to generate a photonic nanojet, which provides the possibility to break the diffraction limit in ORPAM by reducing the size of the excitation volume. In our simulations using microspheres, we observed improvement in the lateral resolution up to ∼fourfold compared to conventional ORPAM. The method is simple, cost effective, and can provide far-field resolution. This approach may provide new opportunities for many biomedical imaging applications that require finer resolution.

1.

Introduction

Photoacoustic imaging (PAI) is a rapidly emerging hybrid biomedical imaging modality to study organelles to organs.13 PAI combines the advantage of rich optical contrast and high ultrasonic resolution at depth. In PAI, a short-pulsed laser source irradiates the tissue. Due to absorption of incident energy by the tissue, there is a local temperature rise, which in turn produces pressure waves emitted in the form of acoustic waves, known as photoacoustic (PA) waves. In photoacoustic tomography (PAT), nonfocused wideband ultrasound detectors noninvasively (placed around the target object) acquire the PA waves generated from the tissue.47 Reconstruction techniques are used to map the initial pressure rise within the tissue from the measured PA waves.812 Initial pressure is in turn related to the absorption coefficient of the tissue.

In photoacoustic microscopy (PAM), typically a focused ultrasound transducer or a focused light source (or both in a confocal geometry) is used and depth-resolved PA signals are recorded [similar to the one-dimensional acoustic signals or A-lines in pulse-echo ultrasound imaging]. Then, a raster scan is done to form a two-dimensional (2-D) or three-dimensional map of optical absorption inside the tissue without the use of any explicit reconstruction algorithms.13 Depending on whether the ultrasound detector or excitation light focus is finer, PAM is classified into acoustic-resolution PAM (ARPAM)14,15 and optical-resolution PAM (ORPAM).1620 The ARPAM, where the detector focus is finer, has achieved a lateral resolution of 45μm and an imaging depth of 3 mm with 50 MHz, 0.44-NA ultrasound transducer.15 The ORPAM, where the optical focus is finer, offers optical diffraction-limited lateral resolution which can be as small as micrometers or sub micrometers. In both ORPAM and ARPAM, the axial resolution is limited by the detector bandwidth (typically 50, 75, 125 MHz).21,22 In this work, we will be discussing only the lateral resolution of ORPAM. The penetration depth of ORPAM is slightly higher than or comparable to the existing optical modalities, such as optical, fluorescence, two-photon microscopy, and optical coherence tomography. However, ORPAM is sensitive to optical absorption with a relative sensitivity of 100% due to background-free detection, whereas the other modalities are sensitive to fluorescence or optical scattering.23,24

The optical-diffraction–limited lateral resolution in ORPAM is given by Ro=0.51λ/NA, where λ denotes the optical wavelength and NA is the numerical aperture of the optical objective.3 Since Ro mainly depends on optical focus, a straight forward method for improving it is to increase the NA. However, the shortcomings of the high-NA objectives (very expensive and short working distances) reduce the flexibility of the system for in vivo imaging. A 220 nm (λ/2) subwavelength (yet diffraction limited) resolution was achieved by using a water-immersion optical objective with NA=1.23 and λ=532nm.16 Thus, in ORPAM, the optical focal spot is the key for controlling the lateral resolution—the tighter the optical focus, the finer the image resolution.

Several recent publications have shown that microspheres could be a relatively simple way of achieving far-field super-resolution in optical microscopy.2529 A resolution of 50 nm was achieved using fused silica microspheres.26 When a microsphere is illuminated by a light beam, a subdiffraction photonic nanojet is generated on the shadow side of the microsphere. The existence of such a nanojet provides the possibility of breaking the diffraction limit in optical microscopy by greatly reducing the size of the excitation volume, and the signal generated in the excitation volume can still be collected in the far-field. Using photonic nanojets, super-resolution imaging was successfully demonstrated in optical microscopy,2628 fluorescence microscopy,30 Raman microscopy,31 and coherent anti-stokes Raman scattering microscopy.32 Not much work has been reported on super-resolution ORPAM imaging until now. Using a double-excitation process, subdiffraction-limited lateral resolution was recently demonstrated in photoimprint PAM.33 A blind-deconvolution optical-resolution photoacoustic microscopy (BD-PAM) that can provide a lateral resolution twofold finer than that of conventional OR-PAM is demonstrated.34 No work has been reported until now on the use of photonic nanojets for super-resolution ORPAM imaging. This approach is simple, cost effective, and can provide far-field super-resolution.

In this work, the possibility to achieve super-resolution in a water medium using microspheres was first investigated. In COMSOL, Gaussian foci were generated using ideal optical objectives with numerical apertures NA=1.0, 0.7, 0.5 (typically used in ORPAM) in a water medium for different incident wavelengths λ1=532nm, λ2=850nm, and λ3=1064nm. Then photonic nanojets were generated by focusing the Gaussian beams through the dielectric microspheres with diameters of (D) 5, 10, and 15μm. The materials used for the microsphere are n2=1.46 (fused silica), 1.6 (polystyrene), and 2.20 (TiO2–BaO–ZnO glass) and they are embedded in media with a refractive index of n1=1.33 (water). The fluence maps generated from COMSOL for Gaussian foci and photonic nanojets were used in a PAM simulation study using the k-wave MATLAB toolbox.35 Super-resolution in ORPAM using microspheres was achieved. In this simulation study, fourfold enhancement in the lateral resolution compared to conventional ORPAM was observed.

2.

Methods

In this work, COMSOL Multiphysics software version 4.2a was used for simulating the Gaussian foci and photonic nanojets in a water medium. A desktop with an Intel xeon 3.4 GHz 64-bit processor, 16-GB RAM with a Windows operating system was used for all the COMSOL simulations done in this study. The fluence maps generated from COMSOL were then used in MATLAB for k-wave simulation. The k-wave toolbox,36 an open source MATLAB toolbox, was used to generate the simulated PA microscopic data. The k-wave toolbox was used only to simulate the PA data generation; no reconstruction technique was used for image formation. A desktop computer with an Intel i5, 2.8 GHz 32-bit processor, 4-GB RAM with a Windows operating system was used for all the k-wave simulations in this study.

2.1.

Simulation of Gaussian Foci/Photonic Nanojets in COMSOL

COMSOL Multiphysics software works based on the finite element method (FEM) for various physics and engineering applications. Based on electromagnetic theory, a 2-D FEM simulation was designed. The simulation model is illustrated in Fig. 1(a). With scattering boundary conditions (black lines), simulations were carried out in a rectangular area (30×60μm2), which includes a water region and a microsphere region. The medium blue region is water, defined by a refractive index n1=1.33. A dielectric microsphere (green region) defined by a refractive index (n2) was embedded in water. The materials used for microspheres in the simulation are fused silica (n2=1.46), polystyrene (n2=1.60), and TiO2–BaO–ZnO glass (n2=2.20). We simulated a focusing laser beam by a TM Gaussian beam propagating along the +y-axis with a full width at half maximum (FWHM) beam waist diameter of (λ/2NA) or an e2 radius of (λ/2.35NA) in intensity,32 where λ (532, 850, or 1064 nm) is incident wavelength, and NA (0.5, 0.7, or 1.0) is the numerical aperture of the objective. The beam waist was always centered laterally in the simulation area (x=0), but the y-position varied for different simulations. In the model settings, a smaller mesh element (20 nm) was specified in the Gaussian focus/photonic nanojet region. Outside this region, a larger mesh element (200 nm) was used. The computational time for calculating a 30×60-μm2 surface plot was 9 s. Several simulations were carried out by changing the diameter of the microsphere (D), the incident wavelength (λ), the numerical aperture of the objective (NA), and the refractive index of the microsphere (n2). Table 1 summarizes the FWHM values of the Gaussian foci and nanojets for various cases. Table 1 is discussed in detail in Sec. 3.1.

Fig. 1

Schematic diagram of the two-dimensional simulation geometry for (a) photonic nanojet generation using finite element method (FEM) in COMSOL, and (b) photoacoustic microscopy imaging using k-wave toolbox in MATLAB. Here, n1 and n2 are the refractive indices of water and microsphere, respectively. UST: ultrasound transducer (point detector). D: diameter of the microsphere. d: distance between the targets.

JBO_19_11_116003_f001.png

Table 1

Full width half maximum (FWHM) of Gaussian focus (fg) and photonic nanojet (fp) for different incident wavelengths (λ) in water (n1=1.33). (a) D=5  μm, NA=1.0; (b) D=10  μm, NA=0.7; (c) D=15  μm, NA=0.5. Here, D: diameter of the microsphere, n2: refractive index of micropsphere, NA: numerical aperture of the light focusing lens. Both the FWHM values and the FWHM divided by the incident wavelength (λ) are given in the table. The symbol ‘X’ in indicates that a nanojet was not observed in these cases.

NAn2λ1=532nmλ2=850nmλ3=1064nm
fgfpfg/fpfgfpfg/fpfgfpfg/fp
(a) D=5μm
1.01.460.64λ1 (343 nm)0.50λ1 (266 nm)1.280.59λ2 (499 nm)0.44λ2 (375 nm)1.340.68λ3 (724 nm)0.44λ3 (464 nm)1.55
1.600.39λ1 (208 nm)1.640.36λ2 (309 nm)1.640.33λ3 (350 nm)2.06
2.20XX0.33λ2 (281 nm)1.790.31λ3 (327 nm)2.19
(b) D=10μm
0.71.460.83λ1 (441 nm)0.64λ1 (348 nm)1.300.78λ2 (662 nm)0.62λ2 (525 nm)1.260.77λ3 (821 nm)0.60λ3 (640 nm)1.28
1.600.52λ1 (274 nm)1.600.61λ2 (522 nm)1.280.45λ3 (481 nm)1.71
2.20XX0.35λ2 (300 nm)2.230.29λ3 (312 nm)2.66
(c) D=15μm
0.51.461.05λ1 (556 nm)0.82λ1 (437 nm)1.281.03λ2 (876 nm)0.80λ2 (684 nm)1.291.09λ3 (1164 nm)0.87λ3 (928 nm)1.17
1.600.65λ1 (345 nm)1.620.66λ2 (559 nm)1.560.71λ3 (756 nm)1.44
2.20
XX0.39λ2 (335 nm)2.640.29λ3 (305 nm)3.82

2.2.

Simulation of Photoacoustic Microscopic B-Scan Images using k-Wave Toolbox in MATLAB

Figure 1(b) shows the simulation geometry used for the simulation of B-scan ORPAM images’ acquisition. For simplicity, all simulations were done in 2-D without losing any generality. A computational grid of 251×1501pixels (the white region of size 5×30μm, 20nm/pixel) was used. A perfectly matched boundary layer to satisfy the boundary condition was used for the generation of the forward PA data. A point detector (the red dot) was used for the simulations and was placed at a distance of 26μm from the target center. The center frequency and the bandwidth of the detector were taken as 125 MHz and 70%, respectively. The transducer parameters are chosen to match the transducers typically used in an ORPAM system,37 however, instead of a large focusing transducer, a point detector was used to simplify the simulations. Other transducers with lower or higher center frequencies (and bandwidths) can also be used in the simulation. The axial resolution in PAM is related to the bandwidth of the ultrasound transducer and will not affect the lateral resolution of the imaging system. The lateral resolution is primarily determined by the focusing of light in the ORPAM imaging system.

The scanning region was restricted to 101×1501pixels (the orange region of size 2×30μm2). For forward PA simulation, the time step chosen was 0.05 ns with a total of 600 time steps. The simulations assumed a sound speed of 1500m/s in the medium. For simplicity, the medium was considered acoustically homogeneous and there was no absorption or dispersion of sound. From the simulated pressure waves, PA waves were down sampled five times to mimic a data acquisition system with 4GigaSamples/s. In all the cases, 1% noise was added with the data, resulting in a 40-dB signal-to-noise ratio. The PAM scan head (in simulation the light source and the detector pair) was moved along the x-direction with a step size of 20 nm, resulting in a total of 101 A-lines per B-scan. The A-line data were stacked after taking Hilbert transform to form a B-scan PAM image. The computational time for each A-line was 2 min 10 s.

Three numerical phantoms were used for the study. The first phantom was a line target with a 0.02-μm width along x-axis and a 1-μm length along y-axis (width expected lateral resolution of the imaging system, making it a line target). This was used as a line object to measure the point spread function along the lateral direction. The second object consisted of two square targets with a side length 1μm separated by a distance (d) of 0.5μm. The third one consisted of two square targets with a side length 1μm separated by a distance (d) of 0.2μm. In all the phantoms, the targets were binary, therefore, outside the targets there was no initial pressure rise and inside the target the initial pressure rise was 1.

3.

Results and Discussion

3.1.

Gaussian Foci and Photonic Nanojets in Water Medium

As discussed in Sec. 2.1, simulations for the incident wavelengths (λ1=532nm, λ2=850nm, and λ3=1064nm), sphere diameters (D=5, 10, 15μm), and numerical apertures (NA=1.0, 0.7, and 0.5) were carried out. The FWHM of Gaussian focus (fg) and photonic nanojet (fp) for different incident wavelengths (λ) in water (n1=1.33) are shown in Table 1. Both the FWHM values and the FWHM divided by the incident wavelength (λ) are shown in this table. The FWHM values are significantly affected by all the parameters n2, NA, D, and λ. From the table, it is clear that the FWHM values of the photonic nanojets are much smaller than the corresponding Gaussian foci. It can be seen that in all cases the FWHM has decreased when n2 is increased (1.46 to 2.2) or NA is increased (0.5 to 1.0) or λ is decreased (1064 to 532 nm). These simulations indicate that it is possible to achieve super-resolution imaging in water with the aid of microspheres. The symbol “X” in the table indicates that a nanojet was not observed in these cases. A few representative fluence maps from COMSOL simulations are shown in Fig. 2. For a better view, a 10×15μm2 area around the Gaussian foci/photonic nanojet is shown here. Figure 2(g) shows the normalized line profiles along with the FWHM values. The fluence maps generated for Gaussian foci and photonic nanojets were then used for k-wave simulation in MATLAB for the ORPAM light excitation source.

Fig. 2

Numerical simulation of Gaussian foci and photonic nanojets using FEM in COMSOL: Gaussian beam foci for (a) λ1=532nm, D=5μm, NA=1.0, (c) λ2=850nm, D=10μm, NA=0.7, (e) λ3=1064nm, D=15μm, NA=0.5, photonic nanojets for (b) λ1=532nm, D=5μm, NA=1.0, n2=1.6, (d) λ2=850nm, D=10μm, NA=0.7, n2=2.2, (f) λ3=1064nm, D=15μm, NA=0.5, n2=2.2. (g) Normalized line scan profiles along the lines A and B indicated on Figs. 2(e) and 2(f), respectively. COMSOL simulations were carried out in 30×60-μm2 area, however, 10×15-μm2 area around the focus is shown here.

JBO_19_11_116003_f002.png

3.2.

Optical Resolution Photoacoustic Microscopic B-Scan Imaging with Photonic Nanojet in k-Wave

First, the line target numerical phantom with a 20-nm width and 1000-nm length was scanned. Figures 3(a) and 3(b) show the 2-D B-scan PAM images of the line target excited with a normal Gaussian focused beam with FWHM 1164 nm [Fig. 2(e)] and excited with a photonic nanojet with FWHM 305 nm [Fig. 2(f)], respectively. It is evident that the lateral resolution has improved with the photonic nanojet excitation. Figure 3(c) shows the normalized lateral line profiles along the lines indicated on Figs. 3(a) and 3(b). From the line profiles, the lateral resolution of the ORPAM system was quantified to be 1135 and 355 nm, respectively (very close to the values measured in COMSOL from the excitation beam profiles). The slight difference is due to the grid size (20 nm) and the nonideal line target (line width 20 nm) used in the k-wave simulation.

Fig. 3

k-wave simulation of B-scan PAM image with a line absorber target object [20-nm line width, length 1000 nm]. B-scan image obtained with (a) Gaussian focus light [shown in Fig. 2(e)] excitation, (b) photonic nanojet [shown in Fig. 2(f)] excitation. (c) Normalized line scan profiles (point spread function) along the dotted lines A and B indicated on 3(a) and 3(b), respectively.

JBO_19_11_116003_f003.png

The second numerical phantom where two square targets with side lengths 1μm and seperated by a distance d=0.5μm were imaged next. Figure 4(a) shows the ORPAM B-scan image of the targets obtained by exciting with normal the Gaussian-focused beam [Fig. 2(e)]. The targets were not resolved in this case beacuse of the poor lateral resolution (the optical spot size is larger than the speration beteen the targets). Whereas in Fig. 4(b) the targets were clearly resolved when ORPAM B-scan images were generated with photonic nanojet excitation [Fig. 2(f)]. Figure 4(c) shows the normalized line scan profiles along the lines indicated on Figs. 4(a) and 4(b). Once again, the line profiles in Fig. 4(c) confirmed the resolving ability of a separation of a 500-nm gap with the photonic nanojet excited ORPAM.

Fig. 4

k-wave simulation of PAM B-scan images of two square (size 1μm) targets separated by a gap of 500 nm. B-scan image with (a) Gaussian focus [shown in Fig. 2(e)] excitation, (b) photonic nanojet [shown in Fig. 2(f)] excitation. (c) Normalized line scan profiles along the dotted lines A and B indicated on 4(a) and 4(b), respectively.

JBO_19_11_116003_f004.png

Next, similar exercises were repeated with a smaller gap of 200 nm between the two square target objects to determine the smallest gap the ORPAM system will be able to resolve. The excitation light source used was the smallest photonic nanojet that was simulated with the set of parameters in this study [FWHM 208 nm Fig. 2(b)]. Figure 5(a) shows the ORPAM B-scan image and Fig. 5(b) shows the normalized profile along the line indicated on Fig. 5(a). The targets are resolved in this case as expected, demonstrating the ability of breaking through the diffraction-limited lateral resolution in ORPAM with the help of photonic nanojets to achieve super-resolution.

Fig. 5

k-wave simulation of PAM B-scan images of two square (size 1μm) targets separated by a gap of 200 nm. (a) B-scan image with photonic nanojet [shown in Fig. 2(d)] excitation. (b) Normalized line scan profile along the dotted line indicated on (a).

JBO_19_11_116003_f005.png

The super-resolution ORPAM modality presented in this work promises to enhance the lateral resolution beyond the diffraction limit. This approach may revolutionize the fundamental biological studies involving cellular structures, vasculature, and microcirculation systems. Melanoma arises from deadly skin cancer melanocytes, and so on.16,38 Currently, the diagnosis of melanoma is done based on invasive biopsy and inaccurate visual inspection. Super-resolution ORPAM (being an in vivo, label-free, noninvasive technique) will help us to diagnose melanoma in the early stages, a key to successful treatment. Tumor angiogenesis, a hallmark of cancer, is presently imaged either in vivo by clinical methods at low resolution or ex vivo by nonclinical methods at high resolution.16,39 Super-resolution ORPAM, having high endogenous contrast, can image angiogenic microvessels in vivo. However, the drawback with this approach is limited penetration depth compared to the conventional ORPAM. The generated nanojets are limited in depth, as observed from the simulations. One can play with various parameters such as the diameter of the microsphere, the refractive-index of the material of the microsphere, etc., to move the focusing spot (nanojets) at various depths from the surface of the microsphere. We can also use some form of z-scanning to increase the depth of imaging. We may not reach the depth penetration of conventional ORPAM, however, it can be significantly improved. In the future, we will work on experimentally implementing the super-resolution ORPAM and applying it to cellular imaging. Improving the penetration depth of super-resolution ORPAM will also be a challenging task to overcome in the future.

4.

Conclusions

In this work, through numerical simulations it has been demonstrated that it is possible to achieve super-resolution in optical resolution PAM using microspheres. The formation of the Gaussian foci and photonic nanojets in a water medium were studied for various parameters of the light focusing objective, microsphere refractive index, size, and excitation wavelengths. For a given NA, by proper selection of the incident wavelength and microsphere refractive index more than twofold lateral resolution improvement is achievable in conventional ORPAM. In our simulations, fourfold improvement in lateral resolution was observed with 0.5 NA, 1064 nm, n2=2.2, D=15μm. This method is simple, cost effective, and can provide far-field super-resolution which allows flexibility of the system for in vivo imaging. This approach opens new opportunities for many biomedical imaging applications, such as cellular, subcellular, microcirculation systems, and tumor angiogenesis that require higher lateral resolution than that obtained at present. In the future, we will experimentally implement the photonic nanojet approach to achieve super-resolution in ORPAM.

Acknowledgments

The work was funded by the University Start-up Grant at Nanyang Technological University, National Natural Science Foundation of China (Grant No. 61301053), and the Fundamental Research Funds for the Central Universities.

References

1. 

X. D. Wanget al., “Noninvasive laser-induced photoacoustic tomography for structural and functional in vivo imaging of the brain,” Nat. Biotechnol., 21 (7), 803 –806 (2003). http://dx.doi.org/10.1038/nbt839 NABIF9 1087-0156 Google Scholar

2. 

S. A. Ermilovet al., “Laser optoacoustic imaging system for detection of breast cancer,” J. Biomed. Opt., 14 (2), 024007 (2009). http://dx.doi.org/10.1117/1.3086616 JBOPFO 1083-3668 Google Scholar

3. 

L. V. WangS. Hu, “Photoacoustic tomography: in vivo imaging from organelles to organs,” Science, 335 (6075), 1458 –1462 (2012). http://dx.doi.org/10.1126/science.1216210 SCIEAS 0036-8075 Google Scholar

4. 

X. D. Wanget al., “Three-dimensional laser-induced photoacoustic tomography of mouse brain with the skin and skull intact,” Opt. Lett., 28 (19), 1739 –1741 (2003). http://dx.doi.org/10.1364/OL.28.001739 OPLEDP 0146-9592 Google Scholar

5. 

X. YangL. V. Wang, “Monkey brain cortex imaging by photoacoustic tomography,” J. Biomed. Opt., 13 (4), 044009 (2008). http://dx.doi.org/10.1117/1.2967907 JBOPFO 1083-3668 Google Scholar

6. 

M. Pramaniket al., “Design and evaluation of a novel breast cancer detection system combining both thermoacoustic (TA) and photoacoustic (PA) tomography,” Med. Phys., 35 (6), 2218 (2008). http://dx.doi.org/10.1118/1.2911157 MPHYA6 0094-2405 Google Scholar

7. 

W. Xiaet al., “Design and evaluation of a laboratory prototype system for 3D photoacoustic full breast tomography,” Biomed. Opt. Express, 4 (11), 2555 –2569 (2013). http://dx.doi.org/10.1364/BOE.4.002555 BOEICL 2156-7085 Google Scholar

8. 

Y. XuM. H. XuL. V. Wang, “Exact frequency-domain reconstruction for thermoacoustic tomography—II: cylindrical geometry,” IEEE Trans. Med. Imaging, 21 (7), 829 –833 (2002). http://dx.doi.org/10.1109/TMI.2002.801171 ITMID4 0278-0062 Google Scholar

9. 

G. Paltaufet al., “Iterative reconstruction algorithm for optoacoustic imaging,” J. Acoust. Soc. Am., 112 (4), 1536 (2002). http://dx.doi.org/10.1121/1.1501898 JASMAN 0001-4966 Google Scholar

10. 

B. T. Coxet al., “Two-dimensional quantitative photoacoustic image reconstruction of absorption distributions in scattering media by use of a simple iterative method,” Appl. Opt., 45 (8), 1866 –1875 (2006). http://dx.doi.org/10.1364/AO.45.001866 APOPAI 0003-6935 Google Scholar

11. 

K. Wanget al., “Investigation of iterative image reconstruction in three-dimensional optoacoustic tomography,” Phys. Med. Biol., 57 (17), 5399 –5423 (2012). http://dx.doi.org/10.1088/0031-9155/57/17/5399 PHMBA7 0031-9155 Google Scholar

12. 

J. Prakashet al., “Basis pursuit deconvolution for improving model-based reconstructed images in photoacoustic tomography,” Biomed. Opt. Express, 5 (5), 1363 (2014). http://dx.doi.org/10.1364/BOE.5.001363 BOEICL 2156-7085 Google Scholar

13. 

J. YaoL. V. Wang, “Sensitivity of photoacoustic microscopy,” Photoacoustics, 2 (2), 87 –101 (2014). http://dx.doi.org/10.1016/j.pacs.2014.04.002 JPHODE 0278-2901 Google Scholar

14. 

K. MaslovG. StoicaL. V. Wang, “In vivo dark-field reflection-mode photoacoustic microscopy,” Opt. Lett., 30 (6), 625 –627 (2005). http://dx.doi.org/10.1364/OL.30.000625 OPLEDP 0146-9592 Google Scholar

15. 

H. F. Zhanget al., “Functional photoacoustic microscopy for high-resolution and noninvasive in vivo imaging,” Nat. Biotechnol., 24 (7), 848 –851 (2006). http://dx.doi.org/10.1038/nbt1220 NABIF9 1087-0156 Google Scholar

16. 

C. ZhangK. MaslovL. V. Wang, “Subwavelength-resolution label-free photoacoustic microscopy of optical absorption in vivo,” Opt. Lett., 35 (19), 3195 –3197 (2010). http://dx.doi.org/10.1364/OL.35.003195 OPLEDP 0146-9592 Google Scholar

17. 

S. HuK. MaslovL. V. Wang, “Second-generation optical-resolution photoacoustic microscopy with improved sensitivity and speed,” Opt. Lett., 36 (7), 1134 –1136 (2011). http://dx.doi.org/10.1364/OL.36.001134 OPLEDP 0146-9592 Google Scholar

18. 

G. LiK. I. MaslovL. V. Wang, “Reflection-mode multifocal optical-resolution photoacoustic microscopy,” J. Biomed. Opt., 18 (3), 030501 (2013). http://dx.doi.org/10.1117/1.JBO.18.3.030501 JBOPFO 1083-3668 Google Scholar

19. 

W. Xinget al., “Integrated optical- and acoustic-resolution photoacoustic microscopy based on an optical fiber bundle,” Opt. Lett., 38 (1), 52 –54 (2013). http://dx.doi.org/10.1364/OL.38.000052 OPLEDP 0146-9592 Google Scholar

20. 

W. Shiet al., “In vivo dynamic process imaging using real-time optical-resolution photoacoustic microscopy,” J. Biomed. Opt., 18 (2), 026001 (2013). http://dx.doi.org/10.1117/1.JBO.18.2.026001 JBOPFO 1083-3668 Google Scholar

21. 

M. XuL. V. Wang, “Analytic explanation of spatial resolution related to bandwidth and detector aperture size in thermoacoustic or photoacoustic reconstruction,” Phys. Rev. E, 67 (5), 056605 (2003). http://dx.doi.org/10.1103/PhysRevE.67.056605 PLEEE8 1063-651X Google Scholar

22. 

M. HaltmeierG. Zangerl, “Spatial resolution in photoacoustic tomography: effects of detector size and detector bandwidth,” Inverse Probl., 26 (12), 125002 (2010). http://dx.doi.org/10.1088/0266-5611/26/12/125002 INPEEY 0266-5611 Google Scholar

23. 

K. Maslovet al., “Optical-resolution photoacoustic microscopy for in vivo imaging of single capillaries,” Opt. Lett., 33 (9), 929 –931 (2008). http://dx.doi.org/10.1364/OL.33.000929 OPLEDP 0146-9592 Google Scholar

24. 

L. V. Wang, “Tutorial on photoacoustic microscopy and computed tomography,” IEEE J. Sel. Top. Quantum Electron., 14 (1), 171 –179 (2008). http://dx.doi.org/10.1109/JSTQE.2007.913398 IJSQEN 1077-260X Google Scholar

25. 

A. Heifetzet al., “Photonic Nanojets,” J. Comput. Theoritical Nanosci., 6 (9), 1979 –1992 (2009). http://dx.doi.org/10.1166/jctn.2009.1254 JCTNAB 1546-1955 Google Scholar

26. 

Z. Wanget al., “Optical virtual imaging at 50 nm lateral resolution with a white-light nanoscope,” Nat. Commun., 2 218 (2011). http://dx.doi.org/10.1038/ncomms1211 NCAOBW 2041-1723 Google Scholar

27. 

X. Haoet al., “Microsphere based microscope with optical super-resolution capability,” Appl. Phys. Lett., 99 (20), 203102 (2011). http://dx.doi.org/10.1063/1.3662010 APPLAB 0003-6951 Google Scholar

28. 

A. Darafshehet al., “Optical super-resolution by high-index liquid-immersed microspheres,” Appl. Phys. Lett., 101 (14), 141128 (2012). http://dx.doi.org/10.1063/1.4757600 APPLAB 0003-6951 Google Scholar

29. 

L. Liet al., “Label-free super-resolution imaging of adenoviruses by submerged microsphere optical nanoscopy,” Light: Sci. Appl., 2 (9), e104 (2013). http://dx.doi.org/10.1038/lsa.2013.60 2047-7538 Google Scholar

30. 

H. Yanget al., “Super-resolution biological microscopy using virtual imaging by a microsphere nanoscope,” Small, 10 (9), 1712 –1718 (2014). http://dx.doi.org/10.1002/smll.v10.9 1613-6829 Google Scholar

31. 

J. Kasimet al., “Near-field Raman imaging using optically trapped dielectric microsphere,” Opt. Express, 16 (11), 7976 –7984 (2008). http://dx.doi.org/10.1364/OE.16.007976 OPEXFF 1094-4087 Google Scholar

32. 

P. K. Upputuriet al., “Super-resolution coherent anti-Stokes Raman scattering microscopy with photonic nanojets,” Opt. Express, 22 (11), 12890 –12899 (2014). http://dx.doi.org/10.1364/OE.22.012890 OPEXFF 1094-4087 Google Scholar

33. 

J. Yaoet al., “Photoimprint photoacoustic microscopy for three-dimensional label-free subdiffraction imaging,” Phys. Rev. Lett., 112 (1), 014302 (2014). http://dx.doi.org/10.1103/PhysRevLett.112.014302 PRLTAO 0031-9007 Google Scholar

34. 

J. Chenet al., “Blind-deconvolution optical-resolution photoacoustic microscopy in vivo,” Opt. Express, 21 (6), 7316 –7327 (2013). http://dx.doi.org/10.1364/OE.21.007316 OPEXFF 1094-4087 Google Scholar

35. 

B. E. TreebyB. T. Cox, “k-Wave: MATLAB toolbox for the simulation and reconstruction of photoacoustic wave fields,” J. Biomed. Opt., 15 (2), 021314 (2010). http://dx.doi.org/10.1117/1.3360308 JBOPFO 1083-3668 Google Scholar

36. 

B. E. TreebyE. Z. ZhangB. T. Cox, “Photoacoustic tomography in absorbing acoustic media using time reversal,” Inverse Probl., 26 (11), 115003 (2010). http://dx.doi.org/10.1088/0266-5611/26/11/115003 INPEEY 0266-5611 Google Scholar

37. 

C. Zhanget al., “In vivo photoacoustic microscopy with 7.6-μm axial resolution using a commercial 125-MHz ultrasonic transducer,” J. Biomed. Opt., 17 (11), 116016 (2012). http://dx.doi.org/10.1117/1.JBO.17.11.116016 JBOPFO 1083-3668 Google Scholar

38. 

V. Gray-SchopferC. WellbrockR. Marais, “Melanoma biology and new targeted therapy,” Nature, 445 (7130), 851 –857 (2007). http://dx.doi.org/10.1038/nature05661 NATUAS 0028-0836 Google Scholar

39. 

D. M. McDonaldP. L. Choyke, “Imaging of angiogenesis: from microscope to clinic,” Nat. Med., 9 (6), 713 –725 (2003). http://dx.doi.org/10.1038/nm0603-713 1078-8956 Google Scholar

Biography

Paul Kumar Upputuri is currently working as a research fellow at Nanyang Technological University. He received a PhD in physics from the Indian Institute of Technology (IIT), Madras, in 2010. He has published around 50 papers in journals and conference proceedings. His research areas are biomedical imaging, coherent Raman microscopy, ultrafast-lasers, optical metrology, and instrumentation. He is a member of SPIE, OSA, Indian Laser Association (ILA), and Optical Society of Indian (OSI).

Zhou-Bin Wen is a MS student at the University of Electronic Science & Technology of China. His research area is mainly in electromagnetic simulation.

Zhe Wu is an associate professor at the National Laboratory for Vacuum Electronics, School of Physical Electronics, University of Electronic Science & Technology of China. His research areas are mainly in electromagnetic simulation, optical and electrical characterization for micro/nano structures and application of microwave energy.

Manojit Pramanik is an assistant professor in the School of Chemical and Biomedical Engineering at Nanyang Technological University, Singapore. He received his PhD in biomedical engineering from Washington University in St. Louis, MO, USA. His research interests include development of photoacoustic/thermoacoustic imaging systems, image reconstruction method, breast cancer imaging, molecular imaging, contrast agent development, and Monte Carlo simulation of light propagation in biological tissue.

© 2014 Society of Photo-Optical Instrumentation Engineers (SPIE) 0091-3286/2014/$25.00 © 2014 SPIE
Paul Kumar Upputuri, Zhuo-Bin Wen, Zhenya Wu, and Manojit Pramanik "Super-resolution photoacoustic microscopy using photonic nanojets: a simulation study," Journal of Biomedical Optics 19(11), 116003 (3 November 2014). https://doi.org/10.1117/1.JBO.19.11.116003
Published: 3 November 2014
Lens.org Logo
CITATIONS
Cited by 35 scholarly publications.
Advertisement
Advertisement
RIGHTS & PERMISSIONS
Get copyright permission  Get copyright permission on Copyright Marketplace
KEYWORDS
Photonic nanostructures

Super resolution

Photoacoustic microscopy

Image resolution

Photonic microstructures

Sensors

Water

Back to Top